Hengxin Ren‡
a,
Xinyu Sun‡a,
Tong Lia,
Zhixin Rena,
Chaoyu Songac,
Yuguang Lv
*ab and
Ying Wang*a
aCollege of Pharmacy, Jiamusi University, Jiamusi 154007, China. E-mail: yuguanglv@163.com
bCollege of Materials Science and Engineering, Jiamusi University, Jiamusi 154007, China
cSchool of Chemistry and Chemical Engineering, Shanghai Jiao Tong University, Shanghai 200000, China
First published on 25th April 2025
Herein, a new Z-scheme BiVO4/Cu2O/PPy heterostructure photocatalyst was developed with bismuth nitrate and ammonium vanadate as the precursors and sodium dodecyl benzyl sulfonate as the soft template. Through the spatial confinement effect of the sodium dodecyl benzyl sulfonate soft template, peanut-like BiVO4 and the BiVO4/Cu2O/PPy heterojunction were synthesized. The best performance was observed for BiVO4/Cu2O/PPy (5%), and its photodegradation rate was 6.57 times higher than that of pure BiVO4. The mechanism study showed that a light hole (h+), superoxide radical (·O2−), and hydroxyl radical (·OH) participated in the CO2 reduction process, which was different from the pure BiVO4 reaction system. Therefore, the proposed approach provides a new method for applying BiVO4/Cu2O/PPy photocatalysts and developing the same type of heterojunction photocatalyst, and they have effective practical application for environmental remediation.
Generally, it is difficult for a single component to simultaneously exhibit a wide light absorption range, effective separation of photogenic carriers, abundant reaction sites and strong REDOX capacity. In order to improve photocatalytic efficiency, heterojunctions, which can promote the separation of photogenerated carriers and integrate the respective advantages of each component, are considered one of the most effective ways. According to the development history, there are three generations of Z-scheme heterostructures (Fig. 1):6–8 (1) traditional Z-scheme heterojunction with a shuttle REDOX medium (Fig. 1a); (2) all-solid-state Z-scheme heterojunction with an electronic medium (Fig. 1b); (3) direct Z-scheme heterojunction with electron-free media (Fig. 1c). The photogenerated electrons and holes of first-generation heterojunction are consumed by the REDOX medium (A/D), which severely weakens photocatalytic activity. In addition, these REDOX media suffer from the drawbacks of shading, pH sensitivity and suitability for liquid media alone, greatly restricting their wide application. Thus, in order to solve the above-mentioned problems of first-generation heterojunctions, the concept of all-solid Z-scheme heterojunction was proposed. In the all-solid-state Z-scheme heterojunction, an electronic medium with good conductivity binds the two semiconductors tightly, thus replacing the shuttle REDOX medium in the first-generation Z-scheme heterojunction. This structure allows the photogenerated electrons and holes to realize spatial separation, thus improving the photocatalytic activity. However, the cost of introducing electronic media, competitive light absorption and demanding structural control requirements limit its application. As a result, the third generation of direct Z-scheme heterostructures emerged without the need for electronic media. Close contact of A and B semiconductors can produce photogenerated electrons and holes under photoexcitation. Electrons in CB of B can directly combine with holes in VB of A. At the same time, the remaining VB holes in B and the electrons in CB of A maintain the initially strong REDOX capacity.
Cuprous oxide (Cu2O) is a reductive semiconductor material with application potential. However, the small Eg of Cu2O (about 2.1 eV) limits the separation efficiency of photogenerated electron–hole pairs, and Cu2O is easily disproportionated to form Cu and CuO during the photocatalytic reaction, thus reducing its lifetime and activity.9 Bismuth vanadate (BiVO4) is a widely studied oxide semiconductor material. As a typical low-cost N-type semiconductor, monoclinic crystal BiVO4 has a unique band edge position suitable for water decomposition, with a long hole diffusion distance (about 70 nm), high carrier lifetime (40 ns), and relatively stable photochemical properties.10 Therefore, it has significant potential in solar energy conversion and environmental purification. However, the high recombination rate of photoinduced carriers, together with the low carrier mobility and reduction potential limit the photocatalytic activity of BiVO4.11–13
Combining the high oxidation potential valence band of BiVO4 with the high reduction potential conduction band of Cu2O to construct the BiVO4/Cu2O heterojunction can effectively separate the electron–hole pair, improve the interface charge transfer efficiency and enhance the photocatalytic effect.14 Kim et al.15 demonstrated that Z-scheme heterostructures have strong reduction and oxidation potentials by constructing a Z-scheme charge flow on a three-dimensional nanowire array structure (BVO/C/Cu2O). The photocatalytic conversion rate of CO2 to CO from BVO/C/Cu2O nanowire arrays is 3.01 mol g−1 h−1, which is 9.4 times and 4.7 times that of Cu2O mesh and Cu2O nanowire arrays, respectively.
Generally, a good electronic medium is needed to reduce the possibility of electron recombination and accelerate the process of electron migration from one semiconductor to another to facilitate charge transfer between heterojunction interfaces. Conjugated polymers such as polyaniline (PANI), polypyrrole (PPy), and polythiophene (PT) are considered to be excellent electrical conductors based on their electrochemical response rather than their structure. PPy is a promising material with magnetic-, optical-, and electronic properties associated with metals or semiconductors. It retains the structure and properties of the polymer, such as ease of processing, flexibility, low toxicity, and adjustable electrical conductivity.16–19 However, its poor mechanical properties limit its general application. Many researchers have been working to improve its performance. The doping strategy significantly increases the conductivity and is an effective scheme to promote charge transfer and separation.20 Therefore, PPy often acts as a conductive matrix to enhance the composites' electrical conductivity and photocatalytic activity.21 In addition, introducing PPy can enhance the dispersion and stability of the composite material in the solution, making it easier to contact the reactants, thus improving the reaction rate and efficiency.22
In this study, the high reduction potential of Cu2O, high oxidation potential of BiVO4 and high conductivity of PPy were combined to construct the Z-scheme photocatalytic heterojunction. The photogenic carrier transfer between Cu2O and BiVO4 is enhanced in the presence of PPy as an electronic medium, with the reducing capacity of Cu2O and the oxidation capacity of BiVO4 maintained. In the presence of PPy as an electronic medium, the photogenic carrier transfer between Cu2O and BiVO4 is enhanced, and the reducing capacity of Cu2O and the oxidation capacity of BiVO4 are maintained, which realizes the efficient photocatalytic reduction of CO2.
![]() | (1) |
The valence band potential and conduction band potential of the sample were calculated using the following empirical formula:
EVB = X + 0.5Eg − Ee | (2) |
ECB = EVB − Eg | (3) |
![]() | ||
Fig. 2 XRD diffraction patterns (a and b) and locally amplified XRD diffraction patterns of the prepared samples (c and d). |
PPy is a typical amorphous polymer. Therefore, sharp diffraction peaks will not be observed in its XRD pattern. A wide reflection at 2θ = 27° reflects the presence of PPy.24 For composites, a characteristic diffraction peak at 2θ = 36.5° indicates successful deposition of Cu2O particles on the BiVO4 surface (Fig. 2c and d).25 However, the diffraction signal of Cu2O is relatively weak due to the effective coating of Cu2O particles on BiVO4 and PPy. In addition, no significant diffraction peaks associated with PPy were detected in the XRD pattern of BiVO4/Cu2O/PPy composites due to the low doping ratio of PPy and its relatively low diffraction intensity (Fig. 2d). In summary, the loading of Cu2O and PPy does not significantly affect the crystal structure of BiVO4.
The microstructure and surface morphology of the sample were clearly demonstrated by SEM (Fig. 3). BiVO4 exhibits a peanut-like morphology (Fig. S1a†), which is attributed to the regulatory role of SDBS as a “soft template” during crystal growth. The surface of these particles is relatively smooth, and the average diameter is 257.69 ± 69.13 nm (Fig. S1d†). SDBS molecules have a hydrophilic head and a hydrophobic tail. This amphiphilic property causes it to form micellar structures in solution and acts as a structural template to guide the growth of nanomaterials to form nanoparticles with predetermined shapes and sizes. Cu2O particles have a regular polyhedral structure with a smooth surface and an average particle size of 74.65 ± 21.68 nm (Fig. S2b and e†). The aggregation of BiVO4 and Cu2O particles indicates that they are highly crystalline. PPy particles gather irregularly to form a morphology similar to cauliflower, with an average particle size of 151.49 ± 32.59 nm (Fig. S2c and f†). Finally, the Cu2O and PPy particles are tightly attached around the BiVO4 particles in the BiVO4/Cu2O/PPy composite (Fig. 3).
The contents of Bi, V, O, Cu, C and N in BiVO4/Cu2O/PPy were basically consistent with the actual added ratio according to EDS analysis (Fig. 4, S2 and S3†). The atomic percentage of Bi and V is close to 1:
1, which is consistent with the atomic composition of BiVO4 (Table 1). The element distribution map (Fig. S3†) shows that Bi, V, O, Cu, C and N are uniformly distributed in a specific region, where the distribution of Bi, V and O is consistent with the position of the peanut-like BiVO4 particles, and the distribution of Cu and O is consistent with the position of the regular polyhedron structure of Cu2O. The peanut-like BiVO4 surface is evenly covered with C and N elements, which is attributed to the presence of PPy. SEM and EDS maps confirmed the formation of BiVO4/Cu2O/PPy composites and the tight binding between them, which is favorable for photocatalytic activity.
Element | Wt (%) | At (%) |
---|---|---|
C | 6.02 | 22.77 |
N | 0.78 | 2.52 |
O | 10.49 | 29.79 |
V | 7.41 | 6.61 |
Cu | 44.05 | 31.51 |
Bi | 31.26 | 6.80 |
The morphological characteristics of BiVO4/Cu2O/PPy nanocomposites were further verified by HRTEM (Fig. 5). Cu2O and PPy surround BiVO4, forming a clear coating structure. The measured lattice fringe spacing of 0.306 nm is consistent with the (121) plane of monocline BiVO4, while the 0.290 nm interval is consistent with the (110) plane of Cu2O (Fig. 5b). These observations suggest that BiVO4, Cu2O, and PPy form an almost perfect interface contact.
![]() | ||
Fig. 5 HRTEM images of BiVO4/Cu2O/PPy nanocomposites: (a) overall coating structure; (b) lattice fringes of BiVO4 and Cu2O. |
The molecular spectral characteristics and chemical group information of the samples were analyzed by FT-IR (Fig. 6). The characteristic peaks at 512 cm−1 and 784 cm−1 are caused by the stretching vibration of Bi–O and the bending vibration of ν3 (VO43−), and the overlap of these absorption peaks results in a wide peak between 700–900 cm−1.26 There is a strong absorption peak at 633 cm−1, which belongs to the Cu–O bond stretching vibration of the Cu2O sample.27 The characteristic peaks at 1323 and 1631 cm−1 are attributed to the stretching vibrations of –CN and C
C in the PPy ring.28 These results fully proved the successful synthesis of BiVO4, Cu2O and PPy monomers. In addition, the characteristic peak of Cu2O at 633 cm−1 overlaps with the wide characteristic peak of BiVO4 at 767 cm−1, which only causes small changes in the doping ratio of different Cu2O samples at 633 cm−1, but it proves that the loading of Cu2O is successful. The triplet heterojunction samples with different PPy doping ratios showed the characteristic peak of PPy at 1631 cm−1, which proved the successful loading of PPy. The stretching vibration of –OH at 3423 cm−1 is attributed to the moisture absorbed by the sample.
The light absorption characteristics of the samples were analyzed by UV-vis DRS. The absorption edges of Cu2O and BiVO4 are about 730 nm and 525 nm (Fig. 7a), respectively. The absorption edges of the composites are dispersed between 550 nm and 590 nm. All samples have light absorption properties ranging from ultraviolet to visible light, and the visible light absorption of the samples is enhanced after the formation of heterogeneous structures. The visible light absorption capacity of the sample increases with higher doping ratios. BiVO4/Cu2O recombination ratios of 10% and 5% in BiVO4/Cu2O/PPy show the highest light absorption intensity compared with similar samples. The utilization efficiency of visible light is improved in the composite samples compared to individual Cu2O and BiVO4 components, owing to the enhanced absorption properties discussed above.
The bandgap energy of BiVO4 and Cu2O was estimated using eqn (1), where α represents the absorption coefficient, ν is the optical frequency, Eg is the bandgap energy, A is a constant, and n depends on the transition characteristics of the semiconductor. The value of n is 1 for direct transition of BiVO4 and Cu2O. Thus, the Eg of BiVO4 and Cu2O are 2.43 and 1.95 eV (Fig. 7b), respectively. The literature shows that the electronegativity of BiVO4 and Cu2O are 6.04 (ref. 29) and 4.84 eV,30 respectively and the ECB and EVB values of BiVO4 are 0.33 eV and 2.75 eV, respectively, calculated according to formula (2) and (3). The ECB and EVB values of Cu2O are −0.64 eV and 1.31 eV, respectively.
The XPS spectra of the surface chemical composition and electronic states showed that the samples consisted of Bi, V, O, C, N, and Cu elements (Fig. S4a†), which is consistent with the EDS results. The high-resolution spectrum of Bi 4f can be divided into two characteristic peaks with binding energies of 159.1 eV and 164.5 eV, attributed to Bi 4f2/7 and Bi 4f2/5, respectively (Fig. 8a), indicating the presence of Bi3+ in the composite. The two prominent peaks at 523.6 and 516 eV (Fig. 8b) belong to V 2p3/2 and V 2p1/2, caused by V5+ spin–orbit splitting. The 529.1 and 532.4 eV binding energies in Fig. 8c are attributed to the composite's lattice oxygen and surface-adsorbed oxygen. The chemical states of bismuth, vanadium, and oxygen elements fully demonstrate the presence of BiVO4 in the composite sample.31 The binding energies of 285 eV, 285.9 eV, and 288.6 eV in Fig. 8e are attributed to CC/C–C, C–N, and C
O bonds (Fig. 8d). The C
O bond of the C 1s peak demonstrates efficient polypyrrole coating on the BiVO4 surface.32 In Fig. 8e, the binding energy of 397.2 eV was attributed to the polypyrrole N–H bond. The chemical states of the element demonstrated the presence of PPy.33 Fig. 8f shows the high-resolution XPS of Cu 2p. Two distinct Cu2O signal peaks at 951.5 eV and 931.50 eV correspond to Cu+ 2p1/2 and Cu 2p3/2, respectively. The signal peaks attributed to Cu2+ were located at 954.2 eV, 941.2 eV, and 933 eV. The presence of these satellite peaks implies the presence of CuO impurities in the composite photocatalyst, which may be caused by the oxidation of Cu2O by oxygen in the air.34 The percentage of Bi and V atoms is close to 1
:
1 (Fig. S4b†), which conforms to the composition of BiVO4 atoms. The XRD, FT-IR, and XPS tests show that BiVO4, Cu2O, and PPy were successfully loaded in the composite, which agrees with the morphology characterization results of SEM and TEM.
![]() | ||
Fig. 8 XPS patterns of BiVO4/Cu2O/PPy samples: Bi 4f (a), V 2p (b), O 1s (c), C 1s (d), N 1s (e), Cu 2p (f). |
![]() | ||
Fig. 9 Photocatalytic curves of CO2 reduction to (a) CO and (b) CH4; bar chart for reduction product generation rate of CO2 for different catalysts (c); and stability test of BiVO4/Cu2O/PPy (5%) (d). |
![]() | ||
Fig. 10 PL spectrum (excitation wavelength: 360 nm) (a) and EIS curve (b) of various photocatalysts. |
The REDOX capacity of the sample was measured by ESR. ·OH and ·O2− were successfully captured on the surface of BiVO4/Cu2O and BiVO4/Cu2O/PPy samples (Fig. 11), indicating that both of these samples have oxidation and reduction capabilities. At the same time, the discovery of ·O2− is strong evidence of the Z-scheme electron transport mechanism. However, the simple BiVO4 sample does not capture the ·O2− because the ECB energy level of BiVO4 is 0.34 eV, which is lower than the formation potential of ·O2− (−0.33 eV) and insufficient to generate ·O2−. The signals of ·OH and ·O2− on the surface of the sample of the ternary composite catalyst are stronger, which indicates that the this ternary heterostructure can more effectively use photogenerated charge for the photocatalytic conversion reaction. These results confirm that the combination of Cu2O and PPy can enhance the REDOX capacity of BiVO4, thereby improving the efficiency of photocatalytic CO2 conversion. These findings have important implications for understanding and designing efficient photocatalysts.
Under visible light, BiVO4, Cu2O, and PPy produce e− and h+ and transfer to the respective CB and VB, respectively. Due to the difference in the Fermi levels of the respective CB and VB, e− will theoretically transfer from PPy to Cu2O and then reach the CB of BiVO4. Similarly, h+ will be transferred from BiVO4 to Cu2O and then to the LUMO of PPy. The ECB potential of BiVO4 (0.34 eV) is insufficient to drive ·O2− generation (−0.33 eV vs. NHE), while the LUMO potential of PPy (2.8 eV vs. NHE) also fails to support ·OH production. These thermodynamic limitations contradict the observed radical signals in ESR and trapping experiments, invalidating the proposed electron transport pathway. Consequently, we propose an alternative mechanism.
All three photocatalytic materials can be excited by visible light to produce e− and h+ (Fig. 12). Subsequently, the photogenic e− on BiVO4 (0.34 eV) CB was rapidly transferred to the HOMO of PPy (1.05 eV). However, h+ in the Cu2O valence band (1.31 eV) can quickly migrate to HOMO and recombine with e−. Moreover, photoelectrons on the CB of Cu2O would reduce CO2, and part of h+ on the VB of B of BiVO4 would readily bind to generate ·OH, oxidizing H2O to produce O2. This electron transfer method fits with the Z-scheme heterostructure, and its unique electron transport channel can shorten the migration distance and time of the photogenerated charge, further delaying the reorganization of e− and h+, and finally inducing significant photocatalytic activity.35
![]() | ||
Fig. 12 Mechanism of photocatalytic reduction of CO2 by BiVO4/Cu2O/PPy nanocomposites under visible light. |
Footnotes |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ra08130g |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |